ODE Notes
ODE Notes
ODE Notes
hjmf@maths.ed.ac.uki
Contents
1. Overview and basic concepts
1.1. Initial value problems
1.2. Systems and phase portraits
1.3. Linear systems with constant coefficients
1.4. Autonomous equations
Problems
2. Ordinary differential equations of higher order
2.1. Second order equations
2.2. Some several variable calculus
2.3. Energy
2.4. Hamiltonian vector fields
2.5. Gradient vector fields
Problems
3. Linear vector fields
3.1. Some linear algebra
3.2. Real eigenvalues
3.3. Complex eigenvalues
3.4. The exponential of a matrix
3.5. The case n = 2
3.6. Inhomogeneous systems
3.7. Higher order equations
Problems
4. Stability of nonlinear systems
4.1. Topology of Rn
4.2. Existence and uniqueness
4.3. Linearisation about equilibrium points
4.4. Stability
4.5. Liapunov stability
4.6. Stability and gradient fields
4.7. Limit cycles and the PoincareBendixson theorem
Problems
5. Rudiments of the theory of distributions
5.1. Test functions
5.2. Distributions
1
3
3
3
4
6
6
9
9
9
10
10
11
11
13
13
14
15
16
16
18
19
20
27
27
28
30
30
31
33
34
34
41
41
44
2001/2002
46
48
50
54
54
58
59
61
64
64
66
68
M341 ODE
or
x0 (t) = ax(t) .
and
x(0) = C .
We have just proven that such initial value problems have a unique
solution.
1.2. Systems and phase portraits. We will spend most of this
course studying systems of ODEs, e.g.,
x01 = a1 x1
x02 = a2 x2 ,
(2)
2001/2002
M341 ODE
A = [aij ] =
..
.
an1 an2 . . . ann
For each x R we define a vector Ax Rn whose i-th coordinate is
n
ai1 x1 + + ain xn ,
so that a matrix A is interpreted as a map A :
x 7 Ax. In this notation we write (3) simply as
x0 = Ax .
Rn
Rn
sending
(4)
2001/2002
aR.
M341 ODE
of indirect use here. Although it will not plot the vector field itself, it
will plot the solutions.
1
1
1
(a) 1
(b) 2
(c) 2
1
0
2
0
0 1
1
0
(d) 1
(e) 1
(f) 1 1
1
0
2
1 1
Problem 1.4. For A as in (a), (b), (c) of Problem 1.3, solve the initial
value problem
x0 = Ax ,
x(0) = (k1 , k2 , k3 ) .
and
x0 = Bx .
(b) Let A =
1
2001/2002
. Find solutions u(t), v(t) of x0 = Ax such
2
that every solution can be expressed in the form u(t) + v(t) for
suitable constants , .
M341 ODE
(5)
and
x0 = y .
(6)
n
X
xi yi = x1 y1 + x2 y2 + + xn yn .
i=1
The norm of x is kxk where kxk2 = hx, xi. If x, y : R Rn are differentiable functions, then we have the following version of the Leibniz
rule:
hx, yi0 = hx0 , yi + hx, y 0 i .
(7)
10
2001/2002
(8)
(9)
(10)
Such a vector field is called hamiltonian. More precisely, a hamiltonian vector field in R2n is one of the form
(x, y) 7 (grady H, gradx H) ,
where
gradx H =
H
H
,...,
x1
xn
and
grady H =
H
H
,...,
y1
yn
,
M341 ODE
11
Rn
Rn .
R2
are
12
2001/2002
M341 ODE
13
x1
a11 . . . a1n
x1
a11 x1 + + a1n xn
..
.
A ... 7 ... . . . ... ... =
.
xn
an1 . . . ann
xn
an1 x1 + + ann xn
Henceforth we will not distinguish between a linear transformation and
its associated matrix. Notice that nn matrices can be added and multiplied by real numbers, so they too form a vector space isomorphic to
Rn2 . In addition, matrices can be multiplied, and matrix multiplication
and composition of linear transformations correspond.
I assume familiarity with the notions of trace and determinant of
a matrix.
A subset E Rn is a (vector) subspace if
(i) x + y E for every x, y E, and
(ii) x E for every x E and R.
The kernel of a linear transformation A : Rn Rn is the subspace
defined by
ker A = {x Rn | Ax = 0} Rn .
Similarly, the image is the subspace defined by
im A = {y Rn | y = Ax, x Rn } Rn .
14
2001/2002
(d) im A = Rn
Let A : Rn Rn be a linear transformation. A nonzero vector x Rn
is called a (real) eigenvector if Ax = x for some real number , which
is called a (real) eigenvalue. The condition that be a real eigenvalue
of A means that the linear transformation I A : Rn Rn is not
invertible. Its kernel is called the -eigenspace of A: it consists of all
eigenvectors of A with eigenvalue together with the 0 vector. The
real eigenvalues of A are precisely the real roots of the characteristic
polynomial of A:
pA() = det(I A) .
A complex root of pA () is called a complex eigenvalue of A.
An n n matrix A = [aij ] is diagonal if aij = 0 for i 6= j, and
it is called diagonalisable (over R) if there exists an invertible n n
(real) matrix S, such that SAS 1 = D, with D diagonal. A sufficient
(but not necessary) condition for A to be diagonalisable, is that its
characteristic polynomial should factorise as
pA () = ( 1 )( 2 ) ( n ) ,
where the i are real and distinct. In other words, this means that
pA () should have n distinct real roots. We will summarise this condition as A has real, distinct eigenvalues. Another sufficient condition
for diagonalisability
is that A be symmetric: aij = aji.
a b
Let A =
be a 2 2 matrix. Its trace and determinant are,
c d
respectively, tr A = a + d and det A = ad bc. The characteristic
polynomial is
pA () = 2 (tr A) + det A .
Therefore the condition for A to have real, distinct roots is the positivity of the discriminant
(tr A)2 4 det A > 0 .
3.2. Real eigenvalues. Let A : Rn Rn , x 7 Ax, be a linear vector
field and consider the associated differential equation: x0 = Ax, where
x : R Rn .
If A is diagonal, A = diag{1 , 2 , . . . , n}, we know from Problem
1.7 that this equation has a unique solution for each choice of initial
value x(0). In fact, the solutions are
xi (t) = xi (0)ei t .
The solution depends continuously on the initial conditions (see Problem 3.5).
Suppose that A is diagonalisable, then there exists some constant
invertible matrix S such that D := SAS 1 is diagonal. Consider the
equation x0 = Ax = S 1 DSx. Define y : R Rn by y = Sx. Then
M341 ODE
15
| |
|
V = v1 v2 vn ;
| |
|
3. and invert to obtain the matrix S = V 1 .
For many problems we will not need the explicit form of S, and only
a knowledge of the eigenvalues of A will suffice.
3.3. Complex eigenvalues. It may happen that a real matrix A has
complex eigenvalues; although they always come in complex conjugate
pairs (see Problem 3.8). In this case it is convenient to think of A as a
linear transformation in a complex vector space.
The canonical example of a complex vector space is Cn , the space of
ordered n-tuples of complex numbers: (z1 , z2 , . . . , zn ).
Let A = [aij ] be a n n matrix. It defines a complex linear transformation Cn Cn as follows:
a11 . . . a1n
z1
a11 z1 + + a1n zn
z1
..
.
A ... 7 ... . . . ... ... =
.
zn
an1 . . . ann
zn
an1 z1 + + ann zn
If A is real, then it preserves the real subspace Rn Cn consisting of
real n-tuples: zi = zi .
A sufficient (but not necessary) condition for a complex n n matrix A to be diagonalisable is that its characteristic polynomial should
factorise into distinct linear factors:
pA () = ( 1 )( 2 ) ( n ) ,
where the i C are distinct.
Suppose that A is diagonalisable, but with eigenvalues which might
be complex. This means that there is an invertible n n complex
matrix S such that SAS 1 = D = diag{1 , . . . , n}. The i are in
general complex, but if A is real, they are either real or come in complex
conjugate pairs (see Problem 3.8).
The equation x0 = Ax, where x : R Rn Cn , can be easily solved
by introducing y : R Cn by x(t) = S 1 y(t), where
yi (t) = yi(0)ei t .
16
2001/2002
X
1 j
A = I + A + 12 A2 +
j!
j=0
x(0) = K Rn
is
x(t) = etA K ,
and there are no other solutions.
Although the solution of a linear ordinary differential equation is
given very explicitly in terms of the matrix exponential, exponentiating
a matrixespecially if it is of suffficiently large rankis not practical
in many situations. A more convenient way to solve a linear equation
is to change basis to bring the matrix to a normal form which can be
easily exponentiated, and then change basis back.
3.5. The case n = 2. Recall the following result from linear algebra.
M341 ODE
17
(III) e
1 0
1 1
18
2001/2002
(11)
(12)
(11)
Then every solution of (11) has the form u(t) + v(t) where v(t) is a
solution of the homogeneous equation
x0 = Ax .
(4)
M341 ODE
19
3.7. Higher order equations. Consider the nth order linear differential equation
s(n) + a1 s(n1) + + an1 s0 + an s = 0 ,
(13)
0
1
0 0
0
0
1
.
.
.
..
.. 0
A = ..
(14)
.
0
0
1
an an1
a1
Proposition 3.5. The characteristic polynomial of the matrix A in
(14) is
pA () = n + a1 n1 + + an .
This results says that we can read off the eigenvalues of the matrix
A directly from the differential equation.
Notice that if s(t) and q(t) solve (13) then so do s(t) + q(t) and ks(t)
where k R. In other words, the solutions of (13) form a vector space.
This is an n-dimensional vector space, since the n initial conditions
s(0), s0 (0),..., s(n1) (0) uniquely determine the solution.
At a conceptual level this can be understood as follows. Let F denote the (infinite-dimensional) vector space of smooth (i.e., infinitely
differentiable) functions s : R R. Let L : F F denote the linear
map
s 7 s(n) + a1 s(n1) + + an1 s0 + an s .
A function s(t) solves equation (13) if and only if it belongs to the
kernel of L, which is a subspace of F.
Higher order inhomogeneous equations
s(n) + a1 s(n1) + + an1 s0 + an s = b ,
for b : R
system
(15)
0
0
B(t) =
... .
b(t)
(11)
20
2001/2002
Problems
(Some of the problems are taken from Hirsch & Smale, Chapters 3,4
and 5.)
Problem 3.1. Solve the following initial value problems:
(a) x0 = x
y 0 = x + 2y
x02 = x1 + x2
x(0) = 0 y(0) = 3
x1 (1) = 1 x2 (1) = 1
(c) x0 = Ax
(d) x0 = Ax
x(0) = (0, b, b)
2
0
0
0
A = 0 1
0
2 3
x(0) = (0, 3)
0
3
A=
1 2
y0 x0
b and c constant.
(16)
M341 ODE
21
y 0 = cx by ,
show that if b2 4c > 0, then (16) has a unique solution x(t) for
every initial condition of the form x(0) = u and x0 (0) = v.
(b) If b2 4c > 0, what assumption about b and c ensures that
limt x(t) = 0 for every solution x(t)?
(c) Sketch the graphs of the three solutions of
x00 3x0 + 2x = 0
for the initial conditions
x(0) = 1
and
x0 (0) = 1, 0, 1 .
(e) = 0 > 0
y0 = x
x02 = 2x1
x(0) = 1 y(0) = 1
x1 (c) = 0 x2 (0) = 2
(c) x0 = y
y 0 = x
x(0) = 1 y(0) = 1
(d) x0 = Ax
x(0) = (3, 9)
1 2
A=
2
1
a b
Problem 3.10. Let A =
and let x(t) be a solution of x0 =
b
a
Ax, not identically zero. Show that the curve x(t) is of the following
form:
(a) a circle if a = 0;
(b) a spiral inward toward (0, 0) if a < 0 and b 6= 0;
(c) a spiral outward away from (0, 0) if a > 0 and b 6= 0.
What effect has the sign of b on the spirals in (b) and (c)? What is the
phase portrait if b = 0?
22
2001/2002
(b) x0 = x + z
y 0 = 2z
y 0 = 3y
z 0 = 2y
z 0 = x z
(a) 0
0
y(0) = 7 .
0
1
0 2
1
0
0 0
15
(b) 1 0 17
0 1
7
(i = 1):
5 6
2 1
2 1
(a)
(b)
(c)
3 4
1
2
1 1
0 1
0 1 2
2 0 0
(d)
1 0
(e) 0 0 3
(f) 0 3 0
0 0 0
0 1 3
0 0
(g) 1 0
i
0
i+1
0
(h)
(i)
0 1
0 i
2
1+i
1 0 0 0
1 0 0 0
(j)
1 0 0 0
1 0 0 0
M341 ODE
23
Problem 3.17. For each matrix T in Problem 3.16 find the eigenvalues of eT .
Problem 3.18. Find an example of two linear transformations A, B
on R2 such that
eA+B 6= eA eB .
Problem 3.19. If AB = BA, prove that eA eB = eB eA and eA B =
BeA .
Problem 3.20. Let a linear transformation A : Rn Rn leave invariant a subspace E Rn (that is, Ax E for all x E). Show that eA
also leaves E invariant.
Problem
3.21. Show that there is no real 2 2 matrix S such that
1
0
S
e =
.
0 4
Problem 3.22. Find the general solution to each of the following systems:
(
(
x0 = 2x y
x0 = 2x y
(b)
(a)
y 0 = 2y
y 0 = x + 2y
(
0
0
x =y
x = 2x
(c)
(d)
y 0 = x 2y
y0 = x
z 0 = y 2z
x = y + z
(e)
y0 = z
z 0 = 0
Problem 3.23. In (a), (b) and (c) of Problem 3.22, find the solutions
satisfying each of the following initial conditions:
(a) x(0) = 1, y(0) = 2;
(b) x(0) = 0, y(0) = 2;
(c) x(0) = 0, y(0) = 0.
Problem 3.24. Let A : Rn Rn be a linear transformation that
leaves a subspace E Rn invariant. Let x : R Rn be a solution of
x0 = Ax. If x(t0 ) E for some t0 R, show that x(t) E for all t R.
Problem 3.25. Prove that if the linear transformation A : Rn Rn
has a real eigenvalue < 0, then the equation x0 = Ax has at least one
nontrivial solution x(t) such that limt x(t) = 0.
Problem 3.26. Let A : R2 R2 be a linear transformation and suppose that x0 = Ax has a nontrivial periodic solution, u(t): this means
that u(t + p) = u(t) for some p > 0. Prove that every solution is
periodic with the same period.
24
2001/2002
k 1
0 1 0
(c)
0 k
0 0
(d) 1
1
0 k
Problem 3.31. Let t : R2 R2 be the flow corresponding to the
equation x0 = Ax. (That is, t 7 t (x) is the solution passing through
x at t = 0.) Fix > 0, and show that is a linear map R2 R2 .
Then show that preserves area if and only if tr A = 0, and that in
this case the origin is neither a sink nor a source.
(Hint: A linear transformation is area-preserving if and only if its
determinant is 1.)
Problem 3.32. Describe in words the phase portraits of x0 = Ax for
the following matrices A:
2 0
2 0
1 0
2 0
(a)
(b)
(c)
(d)
.
0 2
0 1
0 2
1 2
Problem 3.33. Let T be an invertible linear transformation on Rn , n
odd. Show that x0 = T x has a nonperiodic solution.
a b
Problem 3.34. Let Let A =
have nonreal eigenvalues. Show
c d
that b 6= 0 and that the nontrivial solution curves to x0 = Ax are spirals
M341 ODE
25
(b) x0 4x t = 0
(c) x0 = y
y0 = 2 x
(d) x0 = y
(e) x0 = x + y + z
y 0 = 4x + sin 2t
y 0 = 2y + t
z 0 = 2z + sin t
(a) tet
(d)
cos 2t + 2 sin 3t
(e) e
cos 2t
(c)
cos 2t + 3 sin 2t
(f) et + 4
(g) 3t 9
Problem 3.37. Find solutions of the following equations having the
specified initial values:
(a) s00 + 4s = 0; s(0) = 1, s0 (0) = 0.
(b) s00 3s0 6s = 0; s(1) = 0, s0 (1) = 1.
Problem 3.38. For each of the following equations find a basis for
the solutions; that is, find two solutions s1 (t) and s2 (t) such that every
solution has the form s1 (t) + s2 (t) for suitable constants , :
(a) s00 + 3s = 0
(b) s00 3s = 0
(c) s00 s0 6s = 0
(d) s00 + s0 + s = 0
State and prove a generalisation of this result for nth order differential
equations
s(n) + a1 s(n1) + + an s = 0 ,
where the polynomial
n + a1 n1 + + an = 0
has n distinct roots with negative real parts.
26
2001/2002
(b) s00 s = 0
(e) s00 s0 + s = 0
(c) s00 + s0 + s = 0
s0 (0) = 1.
M341 ODE
27
28
2001/2002
(17)
The interval I need not be finite and need not be either open nor
closed: [a, b], [a, b), (a, b], (a, b), (, b], (, b), (a, ) and [a, )
are all possible.
Let U Rn be an open set. A vector field f : U Rn defined on U
is said to be C 1 , if it is continuously differentiable; that is, all the n2
partial derivatives are continuous functions U R.
Theorem 4.2. Let U Rn be open, let f : U Rn be a C 1 vector
field and let x0 U. Then there exists a > 0 and a unique solution
x : (a, a) U
of (17) with x(0) = x0 .
There are two significant differences from the linear case: we may
not be able to take U to be all of Rn and we may not be able to extend
the solution from (a, a) to the whole real line.
To illustrate this second point, consider the vector field f : R R
given by f (x) = 1 + x2 . The solution of x0 = f (x) is
x(t) = tan(t c) ,
where c is some constant. Clearly this solution cannot be extended
beyond |t c| < /2. Such vector fields are said to be incomplete.
The differentiability condition on the vector field is necessary. For
example, consider the vector field f : R R given by f (x) = 3x2/3 .
Then both x(t) = 0 and x(t) = t3 solve x0 = f (x) with x(0) = 0. Thus
there is no unique solution. This does not violate the theorem because
f 0 (x) = 2x1/3 is not continuously differentiable at x = 0.
The proof of this theorem is given in Foundations of Analysis, but
we can sketch the idea. Briefly, suppose x(t) solves the initial value
problem
x0 = f (x)
x(0) = x0 .
(18)
M341 ODE
29
where I is some interval in the real line. Then u solves (19) if and only
if it is a fixed point of the operator P . This suggests the following
iterative scheme (called Picards iteration method). One defines a
sequence x1 , x2 , . . . of functions where
Z t
x1 (t) = x0 +
f (x0 ) ds = x0 + f (x0 )t
0
Z t
x2 (t) = x0 +
f (x1 (s)) ds
0
..
.
xk+1 (t) = x0 +
f (xk (s)) ds .
0
30
2001/2002
M341 ODE
31
x is stable if for every neighbourhood W U of x, there is a neighbourhood W1 W such that every solution to (17) with x(0) W1 is
defined and in W for all t > 0. If W1 can be chosen so that in addition
limt x(t) = x, then x is asymptotically stable.
If x is not stable, it is unstable. This means that there exists one
neighbourhood W of x such that for every neighbourhood W1 W of
x, there is at least one solution x(t) starting at x(0) W1 which does
not lie entirely in W .
An equivalent - definition of (asymptotic) stability is given in Problem 4.4.
Stable equilibria which are not asymptotically stable are sometimes
called neutrally stable.
One should note that limt x(t) = x on its own does not imply
stability. (There are counterexamples, but they are quite involved.)
Let x 7 Ax be a linear vector field on Rn . Then the origin is called
a (linear) sink if all the eigenvalues of A have negative real parts.
More generally, a zero x of a C 1 vector field f : U Rn is called a
(nonlinear) sink if all the eigenvalues of the linearisation Df (
x) have
negative real parts.
A linear sink is asymptotically stable, whereas a centre is stable
but not asymptotically stable. Saddles and sources, for example, are
unstable.
The following theorems tell us to what extent we can trust the stability properties of the linearisation of a nonlinear vector field.
Theorem 4.4. Let f : U Rn be a C 1 vector field defined on an open
subset of Rn and let x U be a sink. Then there is a neighbourhood
W U of x such that if x(0) W then x(t) is defined and in W for
all t > 0 and such that limt x(t) = x.
Theorem 4.5. Let U Rn be open and f : U Rn be a C 1 vector
field. Suppose that x is a stable equilibrium point of the equation (17).
Then no eigenvalue of Df (
x) has positive real part.
Morally speaking, these two theorems say that if the linearised systems is unstable or asymptotically stable, then so will be the nonlinear
system in a small enough neighbourhood of the equilibrium point. If
the linearised system is stable but not asymptotically stable, then we
cannot say anything about the nonlinear system. (See Problem 4.12.)
4.5. Liapunov stability. In those cases where linearisation about an
equilibrium point sheds no light on its stability properties (because the
linearisation is neutrally stable, say) a method due to Liapunov can
help. Throughout this section we will let f : U Rn be a C 1 vector
field defined on an open subset U of Rn , and we will let x U be such
that f (
x) = 0.
32
2001/2002
E(x)
= DE(x)f (x) .
Here the right-hand side is simply the operator DE(x) applied to the
vector f (x). If we let t (x) denote the solution to (17) passing through
x when t = 0, then
d
E(x)
= E(t (x))
dt
t=0
by the chain rule.
Definition 4.6. A Liapunov function for x is a continuous function
E : W R defined on a neighbourhood W U of x, differentiable on
W x, such that
(a) E(
x) = 0 and E(x) > 0 if x 6= x;
(b) E 0 in W x.
If in addition, E satisfies
(c) E < 0 in W x,
then it is said to be a strict Liapunov function for x
.
We can now state the stability theorem of Liapunov.
Theorem 4.7 (Liapunov Stability Theorem). Let x be an equilibrium
point for (17). If there exists a (strict) Liapunov function for x then x
is (asymptotically) stable.
Proof. Let > 0 be so small that the closed -ball about x
lies entirely
in W . Let be the minimum value of E on the boundary of this x) of radius centred at x. From (a) we know that
ball, the sphere S (
> 0. Let
n
o
W1 = x B (
x) | E(x) < .
x) since, by (b),
Then no solution starting inside W1 can meet the S (
E is non-increasing on solution curves. Therefore x is stable.
Now assume that (c) also holds so that E is strictly decreasing on
solution curves in W x. Let x(t) be a solution starting in W1 x and
and consider E(x(t)). Showing that limt E(x(t)) = 0 is equivalent
to showing that limt x(t) = x. Since E(x(t)) is strictly decreasing
and bounded below by 0, L := limt E(x(t)) exists. We claim that
L = 0.
Assume for a contradiction that L > 0 instead. Then by the same
argument as in the first part, we deduce that there is some smaller
sphere S ( < r) such that E(x) < L for all points x inside S .
Since E is continuous, it attains a maximum M in the spherical shell
A,R bounded by S and SR . Because E is negative definite, M is
negative. Now consider any solution curve starting inside Sr at time 0,
M341 ODE
33
E(x(s))ds
E(x(t)) = E(x(0)) +
0
E(x(0)) M t ;
but no matter how small M is, for t large enough the right hand side
will eventually be negative, contradicting the positive-definiteness of
E.
One can picture this theorem in the following way. Near x
, a Liapunov function has level sets which look roughly like ellipsoids containing x. One can interpret the condition that E is decreasing geometrically, as saying that at any point on the level set of E, the vector field
f (x) points to the inside of the ellipsoid. If E is merely non-increasing,
then the vector field may also point tangential to the ellipsoid; but in
either case, once inside such an ellipsoid, a solution curve can never
leave again.
There is also a similar result (which we state without proof) concerning the instability of a critical point.
Theorem 4.8 (Liapunov Instability Theorem). Let E : W R be
a continuous function defined on a neighbourhood W U of x and
differentiable in W x. Then if
(a) E > 0 in W x, and
(b) every closed ball centred at x and contained in W contains a point
x where E(x) > 0,
then x is an unstable equilibrium point.
4.6. Stability and gradient fields. Let U Rn be open and let
V : U R be a C 2 function (twice continuously differentiable). Let
f : U Rn be the associated gradient vector field f (x) = grad V (x),
as discussed in Section 2.5.
It follows from the chain rule that
V (x) = kf (x)k2 0 .
Moreover V (x) = 0 if and only if grad V (x) = 0, so that x is an equilibrium point of the gradient system x0 = grad V (x). This, together
with the observations in Section 2.5, allows us to characterise the gradient flows geometrically.
Theorem 4.9. Consider the gradient dynamical system
x0 = grad V (x) .
At regular points, where grad V (x) 6= 0, the solution curves cross level
surfaces orthogonally. Nonregular points are equilibria of the system.
Isolated minima are asymptotically stable.
34
2001/2002
M341 ODE
(b)
(c)
(d)
(e)
x0
x0
x0
x0
35
= x4/3 , x(0) = 0
= x4/3 , x(0) = 1
= sin x, x(0) = 0
= 1/(2x), x(1) = 1
0 1
0 1
0 0
1
0
1
0
0
0
(d)
(e)
0 1
1
0
0 1
1 0
0
0 1 0
36
2001/2002
R2
whose equa-
M341 ODE
37
(b) x2 y 2 2x + 4y + 5
(c) y sin x
(e) x2 + y 2 z
(f) x2 (x 1) + y 2 (y 2) + z 2
Problem 4.17. Find the type of critical point at the origin of the
following system:
(
x0 = x + y x(x2 + y 2 )
.
y 0 = x + y y(x2 + y 2 )
38
2001/2002
1
4
where 0.
M341 ODE
39
(a) Write the corresponding linear system and show that it has an
isolated critical point at the origin.
(b) Show that the function E(x, y) = x2 + y 2 is a Liapunov function.
Deduce that the origin is stable.
(c) Identify the type of critical point and its stability property for
= 0, 0 < < 2, = 2 and > 2. In particular, show that the
origin is asymptotically stable for > 0.
This problem shows that a given Liapunov function may fail to detect
asymptotic stability. (It can be shown that there exists one which
does.) Moral: Liapunov functions are not unique and knowing that
one exists is not the same thing as finding one!
Problem 4.22. For each of the following systems, show that the origin
is an isolated critical point, find a suitable Liapunov function, and prove
that the origin is asymptotically stable:
(
x0 = 3x3 y
(a) 0
y = x5 2y 3
(
x0 = y 2 + xy 2 x3
(c) 0
y = xy + x2 y y 3
(
x0 = 2x + xy 3
(b) 0
y = x2 y 2 y 3
(
x0 = x3 y + x2 y 3 x5
(d) 0
y = 2x4 6x3 y 2 2y 5
where f is a function on the phase plane which is continuous and continuously differentiable on some disk D about the origin.
(a) Show that the origin is an isolated critical point.
(b) By constructing a Liapunov function or otherwise, show that the
origin is asymptotically stable if f is negative definite on D.
40
2001/2002
Problem 4.25. Discuss the stability of the limit cycles and critical
points of the following systems. (Here r2 = x2 + y 2.)
(
(
x0 = x + y + x(r2 3r + 1)
x0 = y + x sin(1/r)
(a) 0
(b)
y = x + y + y(r2 3r + 1)
y 0 = x + y sin(1/r)
(
(
x0 = x y + x(r3 r 1)
x0 = y + x(sin r)/r
(c) 0
(d)
y = x + y + y(r3 r 1)
y 0 = x + y(sin r)/r
Problem 4.26. Does any of the following differential equations have
limit cycles? Justify your answer.
(a) x00 + x0 + (x0 )5 3x3 = 0
(b) x00 (x2 + 1)x0 + x5 = 0
(c) x00 (x0 )2 1 x2 = 0
Problem 4.27. Prove that the following systems have a limit cycle,
by studying the behaviour of the suggested Liapunov function and
applying the PoincareBendixson theorem.
(
x0 = 2x y 2x3 3xy 2
(a)
.
y 0 = 2x + 4y 4y 3 2x2 y
(Hint: Try E(x, y) = 2x2 + y 2.)
(
x0 = 8x 2y 4x3 2xy 2
(b)
y 0 = x + 4y 2y 3 3x2 y
(Hint: Try E(x, y) = x2 + 2y 2.)
M341 ODE
41
42
2001/2002
Proposition 5.2. The space D of test functions has the following easily proven properties:
1. D is a real vector space; so that if 1 , 2 D and c1 , c2 R, then
c1 1 + c2 2 D.
2. If f is smooth and D, then f D.
3. If D, then 0 D. Hence all the derivatives of a test function
are test functions.
We are only considering real-valued functions of a real variable, but
mutatis mutandis everything we say also holds for complex-valued functions of a real variable.
Definition 5.3. A function f : R R is called (absolutely) integrable if
Z
|f (t)|dt < .
R
for any finite interval [a, b]. A special class of locally integrable functions are the (piecewise) continuous functions.
Test functions can be used to probe other functions.
Proposition 5.4. If f is locally integrable and is a test function,
then the following integral is finite:
Z
Z
f := f (t)(t)dt .
R
|f (t)|dt < ,
M341 ODE
43
As the next result shows, test functions are pretty good probes. In
fact, they can distinguish continuous functions.
Theorem 5.5. Let f, g : R R be continuous functions such that
Z
Z
f = g D .
Then f = g.
Proof. We prove the logically equivalent
if f 6= g, then there
R statement:
R
exists some test function for which f 6= g .
If f 6= g there is some point t0 for which f (t0 ) 6= g(t0 ). Without loss
of generality, let us assume that f (t0 ) > g(t0). By continuity this is
also true in a neighbourhood of that point. That is, there exist > 0
and > 0, such that
f (t) g(t)
for |t t0 | .
whence
f 6=
t0
g .
44
2001/2002
M341 ODE
45
D .
(20)
This distribution
cannot be regular: indeed, if there were a function
R
(t) such that = (0) it would have to satisfy that (t) = 0 for
all t 6= 0; but then such a function could not possibly have a nonzero
integral with any test function. Nevertheless it is not uncommon to
refer to this distribution as the Dirac -function.
Distributions which are not regular are called singular.
Distributions obey properties which are analogous to those obeyed
by the test functions. In fact, dually to Proposition 5.2 we have the
following result.
Proposition 5.13. The space D0 of distributions enjoys the following
properties:
1. D0 is a real vector space. Indeed, if T1 , T2 D0 and c1 , c2 R,
then c1 T1 + c2 T2 defined by
hc1 T1 + c2 T2 , i = c1 hT1 , i + c2 hT2 , i
is a distribution.
2. If f is smooth and T D0 , then f T , defined by
hf T, i = hT, f i
is a distribution.
3. If T D0 , then T 0 defined by
hT 0 , i = hT, 0 i
(21)
is a distribution.
Notice that any test function, being locally integrable, gives rise
to a (regular) distribution. This means that we have a linear map
D D0 which is one-to-one by Theorem 5.5. On the other hand, the
existence of singular distributions means that this map is not injective.
Nevertheless one can approximate (in a sense to be made precise below)
singular distributions by regular ones.
46
2001/2002
D .
where we have used the fact that has compact support. Comparing
with equation (20), we see that is the (distributional) derivative of
the step function:
TH0 = .
(22)
(n)
n
X
=
(1)i f (i) (0) (ni) .
i=0
(23)
M341 ODE
47
0
m (n)
t = (1)m m!
(1)m n! (nm)
(nm)!
m>n
m=n
m<n.
(k)
T , = (1)k T, (k)
T D0
D .
(24)
D ,
(25)
D .
48
2001/2002
The distributional ODE (25) can have two different types of solutions:
Classical solutions. These are regular distributions T = Tx ,
where in addition x is sufficiently differentiable so that L x makes
sense as a function. In this case, = Tf has to be a regular
distribution corresponding to a continuous function f .
Weak solutions. These are either regular distributions T = Tx ,
where x is not sufficiently differentiable for L x to make sense as
a function; or simply singular distributions.
Suppose that the differential operator L is in standard form, so that
an 1. Then it is possible to show that if the inhomogeneous term
in equation (25) is the regular distribution = Tf corresponding to a
continuous function f , then all solutions are regular, with T = Tx for
x(t) a sufficiently differentiable function obeying L x = f as functions.
However, a simple first order equation like
t2 T 0 = 0 ,
which as functions would only have as solution a constant function, has
a three-parameter family of distributional solutions:
T = c1 + c2 TH + c3 ,
where ci are constants and H is the Heaviside step function.
5.4. Greens functions. Solving a linear differential equation is not
unlike inverting a matrix, albeit an infinite-dimensional one. Indeed, a
linear differential operator L is simply a linear transformation in some
infinite-dimensional vector space: the vector space of distributions in
the case of an equation of the form (25). If this equation were a linear
equation in a finite-dimensional vector space, the solution would be
obtained by inverting the operator L, now realised as a matrix. In this
section we take the first steps towards making this analogy precise. We
will introduce the analogue of the inverse for L (the Greens function
of L) and the analogue of matrix multiplication (convolution).
Let L be an n-th order linear differential operator in standard form:
L = Dn + an1 Dn1 + + a1 D + a0 ,
where ai are smooth functions.
Definition 5.15. By a fundamental solution for L we mean a distribution T satisfying
LT = .
Fundamental solutions are not unique, since one can add to T anything in the kernel of L. For example, we saw in (22) that (the regular
distribution defined by) the Heaviside step function is a fundamental
M341 ODE
49
for some c R.
One way to resolve this ambiguity is to impose boundary conditions.
An important class of boundary conditions is the following.
Definition 5.16. A (causal) Greens function for the operator L
is a fundamental solution G which in addition obeys
hG, i = 0 D such that supp (, 0) .
In other words, the Greens function G is zero on any test function
(t) vanishing for non-negative values of t. With a slight abuse of
notation, and thinking of G as a function, we can say that G(t) = 0
for t < 0.
As an example, consider the Greens function for the differential
operator L = Dk , which is given by
tk1
G(t) =
H(t) .
(k 1)!
Let us consider the Greens function for the linear operator L above,
where the ai are real constants. By definition the Greens function G(t)
obeys L G = and G(t) = 0 for t < 0. This last condition suggests
that we try G(t) = x(t)H(t), where x(t) is a function to be determined
and H is the Heaviside step function. Computing L G we find, after
quite a little bit of algebra,
"
#
n1 n`1
X
X k + `
ak+`+1 x(k) (0) (`) ,
L G = (L x) H +
`
`=0
k=0
x(n1) (0) = 1 .
We know from our treatment of linear vector fields that this initial
value problem has a unique solution. Therefore the Greens function
for L exists and is unique.
The Greens function is the analogue of an inverse of the differential
operator. In fact, it is only a right-inverse: it is a common feature
of infinite-dimensional vector spaces that linear transformations may
have left- or right-inverses but not both. This statement is lent further
credibility by the fact there is a product relative to which the Dirac
is the identity. This is the analogue of matrix multiplication: the convolution product. Convolutions are treated in more detail in Problems
50
2001/2002
This definition embodies the principle of causality. If the above equation describes the response of a physical system to an external input
f (t), then one expects that the response of the system at any given
time should not depend on the future behaviour of the input.
Problems
Problem 5.1. Let f : R R be an absolutely integrable function of
unit area; that is,
Z
Z
|f (t)|dt <
and
f (t)dt = 1 .
R
M341 ODE
51
hTfn , i (0) =
estimate the integral and show that it goes to zero for large n.)
Problem 5.2. Let f be a continuous function whose derivative f 0 is
also continuous (i.e., f is of class C 1 ). Let Tf denote the corresponding
regular distribution. Prove that Tf0 = Tf 0 .
Problem 5.3. Let a R and define the shifted step function Ha (t) by
(
1, ta
Ha (t) =
0, t<a,
and let THa be the corresponding regular distribution. Prove that
TH0 a = a , where a is defined by
ha , i = (a) D .
Problem 5.4. Let f be a smooth function and T be a distribution.
Then f T is a distribution, as was shown in lecture. Prove that
(f T )0 = f 0 T + f T 0 .
Problem 5.5. Let f be a smooth function. Prove that
n
X
(n)
j n
(1)
f (j) (0) (nj) .
f =
j
j=0
As a corollary, prove that
n<m
0 ,
m (n)
m
n=m
t = (1) m! ,
n
X
n nj (j)
=
.
j
j=0
|t| < 1
t ,
f (t) = 1 ,
t1
1 , t 1 .
52
2001/2002
D2 k 2 ek|t| = 2k .
Use this result to find the causal Greens function for the linear operator
L = D2 k 2 and find a solution x(t) of the inhomogeneous equation
Lx(t) = f (t), where
(
t for 0 < t < 1,
f (t) =
0 otherwise.
Problem 5.9. Find the Greens function for the linear second order
differential operator
L = D2 + aD + b ,
where a, b R. Distinguish between the cases a2 < 4b, a2 = 4b and
a2 > 4b and write the Greens function explicitly in this case. Use this
to solve the inhomogeneous initial value problem
x00 (t) + x0 (t) + x(t) = f (t)
x(t), x0 (t) 0 as t ,
where f (t) is the piecewise continuous function f (t) = H(t) H(t 1).
Sketch (or ask Maple to sketch) x(t) as a function of t. From the sketch
or otherwise, is x(t) smooth?
Problem 5.10. Let , and be test functions. Define their convolution ? by
Z
( ? )(t) := (t s) (s) ds .
R
(a) Show that if has support [a, b] and has support [c, d], then
? has support [a + c, b + d].
(b) Show that ( ? )0 = 0 ? = ? 0 .
(c) Conclude that ? is a test function.
(d) Show that the convolution product is commutative: ? = ? ,
and associative
( ? ) ? = ? ( ? ) .
The following problems get a little deeper into the notion of a distribution. They are not examinable, but some of you might find them
interesting.
Problem 5.11. Let : D D be a continuous linear map; that is,
1. (c1 1 + c2 2 ) = c1 (1 ) + c2 (2 ) for ci R and i D; and
2. if m 0 then (m ) 0 in D.
M341 ODE
53
D ,
(26)
maps D0 to D0 .
Let a, b R with a 6= 0 and define the following operations on functions:
(a ) (t) = (ax) and
(b ) (t) = (t b) .
(b) Prove that a and b map test functions to test functions, and
that they are linear and continuous.
Let a : D0 D0 and b : D0 D0 be their adjoints, defined by (26).
(c) If f is a locally integrable function and Tf the corresponding regular distribution, show that
a Tf = Ta f
and b Tf = Tb f ,
where
(a f ) (t) =
1
f (t/a) and
|a|
(b f ) (t) = f (t + b) .
54
2001/2002
solution: f
6inverse
transform
transform
solution: F
The two most important integral transforms are the Fourier transform (cf. PDE) and the Laplace transform. Whereas the Fourier transform is useful in boundary value problems, the Laplace transform is
useful in solving initial value problems. As this is the main topic of
this course, we will concentrate solely on the Laplace transform.
6.1. Definition and basic properties.
Definition 6.1. Let f (t) be a function. Its Laplace transform is
defined by
Z
L {f } (s) :=
f (t) est dt ,
(27)
0
provided that the integral exists. One often uses the shorthand F (s)
for the Laplace transform L {f } (s) of f (t).
Remark 6.2. The following should be kept in mind:
M341 ODE
55
t .
M et e Re(s)t dt .
56
2001/2002
Transform
Conditions
f (t)
F (s)
convergence
eat
1
sa
cos t
s
s2 + 2
sin t
s2 + 2
cosh t
s
s2 2
sinh t
s2 2
tn
n!
sn+1
eat f (t)
F (s a)
convergence
tn f (t)
f (t)
es F (s)
n1
sn F (s)
f (n) (t)
k=0
t
f (t ) g( ) d
F (s) G(s)
M341 ODE
57
u=0
u=0
t=u
for any function k(t, u) for which the integrals exist. Therefore,
Z Z
st
L {f ? g} (s) =
e f (t u) dt g(u) du
0
u
Z Z
s(u+v)
=
e
f (v) dv g(u) du
(v = t u)
0
0
Z
= L {f } (s)
esu g(u) du
0
58
2001/2002
6.2. Application: solving linear ODEs. As explained in the introduction to this section, the usefulness of the Laplace transform is based
on its ability to turn initial value problems into algebraic equations. We
now discuss this method in more detail.
Let L be a linear differential operator in standard form:
n
X
n
n1
L = D + an1 D
+ . . . a1 D + a0 =
ai Di ,
i=0
We can solve this using the Laplace transform in three easy steps:
1. We take the Laplace transform of the equation:
L {Lx} (s) = L {f } (s) = F (s) .
Using linearity and the expression for the Laplace transform for
x(i) in Table 1 one finds
!
n
X
L {Lx} (s) =
ai si X(s) P (s) ,
i=0
n1
X
pi si
i=0
where pi =
ni1
X
ai+j+1 cj ,
(28)
j=0
to obtain
X(s) = T (s) (F (s) + P (s)) ,
where
T (s) =
n
X
i=0
!1
ai si
1
.
sn + an1 sn1 + + a1 s + a0
M341 ODE
59
which looks like hTf , e i except for the minor detail of the lower limit
of integration and the fact that est is not a test function, since it does
not have compact support. However, for Re(s) > 0, the function est
1
The same cannot be said about the term T (s)P (s), since P (s), being a polynomial, cannot be the Laplace of any function; although it can be shown to be the
Laplace transform of a singular distribution.
60
2001/2002
obeys the next best thing: it decays very fast. The following definition
makes this notion precise.
Definition 6.9. We say that a function f is of fast decay if for all
non-negative integers k, p there exists some positive real number Mk,p
such that
(1 + t2 )p |f (k) (t)| Mk,p
t R .
t R .
uniformly in t.
This notion of convergence agrees with the one for test functions. In
other words, a sequence of test functions converging to zero in D also
converge to zero in S. (See Problem 6.9.)
Recall that in Section 5.2 we defined a distribution to be a continuous linear functional on D. Some distributions will extend to linear
functionals on S.
Definition 6.12. A tempered distribution is a continuous linear
functional on S. The space of tempered distributions is denoted S0 .
This means that T S0 associates a number hT, f i with every f S,
in such a way that if f, g S and a, b are constants, then:
hT, a f + b gi = a hT, f i + b hT, gi .
Continuity means that if fn 0 then the numbers hT, fn i 0 as well.
Proposition 6.13. The space of tempered distributions is a vector
space. In fact, it is a vector subspace of D0 : S0 D0 .
Tempered distributions inherit the notion of weak convergence of
distributions in Definition 5.14.
Not all distributions have a Laplace transform. Let T be a distribution with the property that T (t) = 0 for t < 0. In other words,
M341 ODE
61
L {T } (s) := T, est ,
which then exist for Re(s) > .
Let us compute the Laplace transforms of a few singular distributions:
First, we start with T = :
st
st
L {} (s) = , e
=e =1.
t=0
Notice that this is consistent with the fact that is the identity under
convolution (cf. Problem 5.13). Indeed, let f be a function with Laplace
transform F (s). Then,
L { ? f } (s) = L {} (s)F (s) = F (s) = L {f } (s) .
We generalise to T = (k) :
(k) st
(k)
dk st
k
k st
L
(s) = , e
= (1) , k e
= s e = sk .
dt
t=0
These two results allows us to invert the Laplace transform of any
polynomial.
Finally consider for > 0, defined by h , i = ( ). Its Laplace
transform is given by
st
st
L { } (s) = , e
=e
= es .
t=
Problems
Problem 6.1. Provide all the missing proofs of the results in Table 1.
Problem 6.2. Suppose that f (t) is a periodic function with period T ;
that is, f (t + T ) = f (t) for all t. Prove that the Laplace transform
F (s) of f (t) is given by
Z T
1
F (s) =
f (t) est dt ,
1 esT 0
which converges for Re(s) > 0.
Use this to compute the Laplace transform of the function f (t) which
is 1 for t between 0 and 1, 2 and 3, 4 and 5, etcetera and 0 otherwise.
62
2001/2002
(Hint: Take the Laplace transform of the equation, solve for the Laplace
transform F (s) of f (t), and finally invert the transform.)
Problem 6.4. Show that the differential equation:
f 00 (t) + 2 f (t) = u(t),
subject to the initial conditions f (0) = f 0 (0) = 0, has
Z
1 t
f (t) =
u( ) sin (t ) d
0
as its solution.
(Hint: Take the Laplace transform of both sides of the equation, solve
for the Laplace transform of f and invert.)
Problem 6.5. Suppose that we consider the Laplace transform of tz ,
where z is a complex number with Re(z) > 0. This is given in terms
of the Euler function, defined by
Z
(z) :=
tz1 et dt .
0
Prove that
(z + 1)
,
sz+1
provided Re(z) > 0. (This is required for convergence of the integral.)
Prove that
L {tz } (s) =
(z + 1) = z (z) ,
for Re(z) > 0. Compute ( 12 ).
Problem 6.6. Compute the Laplace transforms of the following functions:
1
(a) f (t) = 3 cos 2t 8e2t
(b) f (t) =
t
(
1 , for t < 1, and
(c) f (t) =
(d) f (t) = (sin t)2
0 , for t 1.
0 , for t < 1,
(e) f (t) = 1 , for 1 t 2, and
0 , for t > 2.
Make sure to specify as part of your answer the values of s for which
the Laplace transform is valid.
M341 ODE
63
5
+ 6f (t) = 0 ,
f (0) = 1, f 0 (0) = 1 ,
2
dt
dt
d2 f (t) df (t)
(b)
2f (t) = et sin 2t ,
f (0) = f 0 (0) = 0 ,
dt2
dt
0 , for 0 t < 3,
d2 f (t)
df (t)
+ 2f (t) = 1 , for 3 t 6, and
(c)
3
dt2
dt
0 , for t > 6,
f (0) = f 0 (0) = 0 .
Problem 6.9. Prove that the space S of functions of fast decay is a
vector space. Moreover show that vector addition and scalar multiplication are continuous operations; that is, show that if fn , gn 0 are
sequences of functions of fast decay converging to zero, then fn +gn 0
and cfn 0 for all scalars c. (This makes S into a topological vector
space.)
Prove as well that convergence in D and in S are compatible. (This
makes D is closed subspace of S.)
64
2001/2002
X
an (t t0 )n = a0 + a1 (t t0 ) + a2 (t t0 )2 + ,
n=0
an tn = a0 + a1 t + a2 t2 + .
(29)
n=0
N
X
an tn
exists.
n=0
M341 ODE
65
X
an tn .
f (t) =
n=0
1 (n)
f (0)
n!
X
an
f (t)dt =
tn+1
tn+1
if R < t1 < t2 < R.
2
1
n+1
t1
n=0
Convergent power series can be added and multiplied, the resulting
power series having as radius of convergence at least as large as the
smallest of the radii of convergence of the original power series. The
formula for the product deserves special attention:
!
!
X
X
X
n
n
an t
bn t
=
cn tn ,
n=0
n=0
n=0
where
cn =
n
X
`=0
a` bn` =
n
X
an` b` .
`=0
66
2001/2002
X
an (t t0 )n
(30)
f (t) =
n=0
af (t) + bg(t)
f (t)/g(t)
subject to
x(i) (t0 ) = ci
(31)
M341 ODE
67
pn tn
and
q(t) =
n=0
qn tn .
n=0
Let R denote the smallest of the radii of convergence of these two series.
If an analytic solution exists, it can be approximated by a power series
x(t) =
cn tn ,
n=0
which converges in some interval around t = 0 and hence can be differentiated termwise. Doing so, we find
X
x (t) =
(n + 1)cn+1 tn
0
X
and x (t) =
(n + 1)(n + 2)cn+2 tn .
00
n=0
n=0
Inserting these power series into the differential equation (31), and using the Cauchy product formulae, we can derive a recurrence relation
for the coefficients cn :
X
1
((` + 1)c`+1 pn` + c` qn` ) ,
(n + 1)(n + 2) `=0
n
cn+2 =
68
2001/2002
Problems
Problem 7.1. Let f (t) and g(t) be analytic at t = 0, with power series
expansions
f (t) =
an t
and g(t) =
n=0
bn tn .
n=0
cn tn .
n=0
Write a recurrence relation for the coefficients cn and solve for the first
three coefficients c0 , c1 , c2 in terms of the coefficients {an , bn }.
Problem 7.2. Let f (t) be analytic at t = 0, with power series expansion
X
f (t) =
an tn .
n=0
(t) =
bn (t a0 )n .
n=0
M341 ODE
69
(32)
where p is a constant.
(a) Is t = 0 an ordinary point? Why? What is the radius of convergence of analytic solutions of equation (32)?
(b) Derive a recurrence relation for the coefficients of power series
solutions of (32) at t = 0.
(c) Show that when p is a non-negative integer, (32) admits a polynomial solution. More concretely show that when p = 0, 2, 4, . . .
(32) admits a solution which is an even polynomial of degree p;
and that when p = 1, 3, 5, . . . it admits a solution which is an odd
polynomial of degree p.
(d) Define the Chebyshev polynomials Tp , as the polynomial found
above, normalised so that Tp (0) = 1 for p even and Tp0 (0) = 1 for
p odd. Calculate the first six Chebyshev polynomials: T0 , T1 , ...,
T5 using the recurrence relation found above.
(Note: The standard Chebyshev polynomials in the literature are
normalised in a different way.)
(e) Show that Chebyshev polynomials are orthogonal with respect to
the following inner product on the space of polynomials:
Z 1
1
hf, gi :=
f (t)g(t) dt ,
1 t2
1
using the following method:
(i) Show that the operator T, defined on a polynomial function
f (t) as Tf (t) := tf 0 (t) (1 t2 )f 00 (t), is self-adjoint relative
to the above inner product.
(ii) Show that Tp is an eigenfunction of T with eigenvalue p2 :
TTp = p2 Hp .
(iii) Deduce that if p 6= q, then hTp , Tq i = 0.